Dynamic Theory Notes
Dynamic Theory Notes
Björn Birnir
Center for Complex and Nonlinear Dynamics
and Department of Mathematics
University of California
Santa Barbara1
1 Introduction 9
1.1 The 3 Body Problem . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2 Nonlinear Dynamical Systems Theory . . . . . . . . . . . . . . . 11
1.3 The Nonlinear Pendulum . . . . . . . . . . . . . . . . . . . . . . 11
1.4 The Homoclinic Tangle . . . . . . . . . . . . . . . . . . . . . . . 18
4 Invariant Manifolds 65
5 Chaotic Dynamics 75
5.1 Maps and Diffeomorphisms . . . . . . . . . . . . . . . . . . . . . 75
5.2 Classification of Flows and Maps . . . . . . . . . . . . . . . . . . 81
5.3 Horseshoe Maps and Symbolic Dynamics . . . . . . . . . . . . . 84
5.4 The Smale-Birkhoff Homoclinic Theorem . . . . . . . . . . . . . 95
5.5 The Melnikov Method . . . . . . . . . . . . . . . . . . . . . . . 96
5.6 Transient Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 99
3
4 CONTENTS
1.1 The rotation of the Sun and Jupiter in a plane around a common
center of mass and the motion of a comet perpendicular to the plane. 10
1.2 The nonlinear pendulum is your typical grandfather’s clock,
where the pendulum makes the angle θ with the vertical line. The
distance that the pendulum travels is the arc length `θ where ` is
the length of the arm and gravity pulls on the pendulum with the
force F1 = mg where m is the mass of the pendulum and g the
gravitational acceleration. . . . . . . . . . . . . . . . . . . . . . . 13
1.3 W u denotes the unstable and W s the stable manifolds, Eu is the
(linear) unstable subspace and Es the (linear) stable subspace. . . . 15
1.4 The phase portrait for the nonlinear pendulum shows four differ-
ent type of solutions. The first type are the stationary solutions
at the origin, that is a center and the one at (±π, 0) that is a sad-
dle. Around the origin there are periodic orbits corresponding to
small oscillations of the pendulum that are called librations. Then
there are two homoclinic connections connecting the saddle to it-
self, because (±π, 0) are really the same point corresponding to
the vertical position of the pendulum. The fourth type of solu-
tions are rotations, above and below the homoclinic connections.
They correspond to rotations of the pendulum around the circle
with increasing velocity as the distance from the origin increases. 17
1.5 The phase portrait of the damped pendulum is similar to the un-
damped one except that the origin has changed to a sink and all
solutions, except the saddle and the two rotations that connect to
the stable manifolds of the saddle, spiral into the sink. . . . . . . . 19
5
8 LIST OF FIGURES
Chapter 1
Introduction
9
10 CHAPTER 1. INTRODUCTION
Figure 1.1: The rotation of the Sun and Jupiter in a plane around a common center
of mass and the motion of a comet perpendicular to the plane.
1.2. NONLINEAR DYNAMICAL SYSTEMS THEORY 11
The Butterfly Effect: A monarch butterfly fluttering its wings in Santa Bar-
bara, California, today may cause a storm in the town of Akureyri, on the north
coast of Iceland, in two weeks time.
What happens is that the instability leads to a loss of information and this is
the perfect nightmare of someone who is trying to use a computer to obtain his
answers. The tiny round-off errors that one finds in any numerical computation
by a computer are being magnified exponentially in time. This means that after a
short time the error in the computation is overwhelming the answer and the result
of the computation is numerical junk.
We will in later chapters associated chaotic behavior with positive Lyapunov
exponents and another way of saying the above is that positive Lyapunov expo-
nents lead to limits on predictability.
direction. The force F2 in Figure 1.3 is balanced by the tension in the arm of the
pendulum. Thus the above equation becomes
m`θ̈ = −mg sin(θ)
Dividing by m` we get that
g
θ̈ + sin(θ) = 0
`
where r
g
ω=
`
is the frequency of the nonlinear pendulum. The energy of the nonlinear pendulum
is obtained by multiplying the equation by θ̇
θ̇θ̈ + ω2 sin(θ)θ̇ = 0
and integrating with respect to t
θ̇2
+ ω2 (1 − cos(θ)) = C
2
where C is a constant. If we damp and drive the nonlinear pendulum its motion is
described by the equation
θ̈ + δθ̇ + ω2 sin(θ) = ε cos(Ωt)
Now we let x = θ and set ω = 1, then the nonlinear pendulum (without damp-
ing and driving) is described by the equation
(1.1) ẍ + sin (x) = 0
x(0) = x0 , ẋ(0) = ẋ0 .
x0 and ẋ0 are the initial position (angle) and initial (angular) velocity of the pendu-
lum. With these initial conditions specified, the initial value problem (IVP) (1.1)
determines the solution x(t) for all time. We will now give a complete qualitative
analysis of the solutions of the Equation 1.1. This is done by the following steps.
1. Write the equations as a first order system.
We let y = ẋ and rewrite the equation as a first order system
d x y
= .
dt y − sin (x)
1.3. THE NONLINEAR PENDULUM 13
Figure 1.2: The nonlinear pendulum is your typical grandfather’s clock, where the
pendulum makes the angle θ with the vertical line. The distance that the pendulum
travels is the arc length `θ where ` is the length of the arm and gravity pulls on
the pendulum with the force F1 = mg where m is the mass of the pendulum and g
the gravitational acceleration.
14 CHAPTER 1. INTRODUCTION
ẍ + sin (x) = 0.
∂x
We linearize the system and let z = ∂x o
, the derivative of x with respect to
the initial condition xo . Then the linearized system becomes
ż = D(x,y) f (x)z
The eigenvalues of this matrix are λ = ±i, with no real eigenvectors. This
says that the stationary solution (x, y) = (0, 0) is marginally stable. Next we
1.3. THE NONLINEAR PENDULUM 15
Figure 1.3: W u denotes the unstable and W s the stable manifolds, Eu is the (linear)
unstable subspace and Es the (linear) stable subspace.
16 CHAPTER 1. INTRODUCTION
y2
E(x, y) = + 1 − cos (x) = constant.
2
This is obviously the same expression as we obtained above with θ = x, θ̇ = y and
ω = 1. The first part of this expression is the kinetic energy and the second part
the potential energy. E is constant for each solution (orbit) of the IVP, and it can
be used to prove the existence of the elliptical orbits around the origin in Figure
4. In this case, since the ODE is two-dimensional and has an integral, we can
1.3. THE NONLINEAR PENDULUM 17
Figure 1.4: The phase portrait for the nonlinear pendulum shows four different
type of solutions. The first type are the stationary solutions at the origin, that is a
center and the one at (±π, 0) that is a saddle. Around the origin there are periodic
orbits corresponding to small oscillations of the pendulum that are called libra-
tions. Then there are two homoclinic connections connecting the saddle to itself,
because (±π, 0) are really the same point corresponding to the vertical position
of the pendulum. The fourth type of solutions are rotations, above and below the
homoclinic connections. They correspond to rotations of the pendulum around
the circle with increasing velocity as the distance from the origin increases.
18 CHAPTER 1. INTRODUCTION
find W u and W s explicitly. The homoclinic orbits approach the stationary solution
(x, y) = (π (mod 2π), 0) as t → ∞ so by continuity they must have the same energy
as this stationary solution. We therefore set the energy equal to the energy of the
stationary solution
y2
+ 1 − cos (x) = E(π, 0) = 2.
2
This gives a first order equation for x that is solved, see Appendix A, to give
the − sign gives W u (π, 0) and the + sign gives W s (π, 0). Notice that both of these
manifolds are globally defined.
Adding damping the pendulum equation (1.1) gives the initial value problem:
The phase portrait of its solutions is shown in Figure 1.3. The stability of the
stationary solution (x, y) = (π (mod 2π), 0) remains the same, this is because it
is hyperbolic and therefore structurally stable. The stationary solution (x, y) =
(0, 0) on the other hand turns into a sink. The homoclinic connections are not
structurally stable and break under the perturbation but one rotation from each
direction connects with the stable manifolds of the stationary solution (x, y) =
(π (mod 2π), 0). Every other solution eventually spirals into the sink, see Figure
1.3.
is more involved. The first problem is that the equation is non-autonomous and
one must analyze the extended phase space, which is three-dimensional, instead
of the two-dimensional phase space above. This is harder but more interesting
things can happen in three dimensions. Poincaré got around this problem by in-
venting the Poincaré map, see Figure 1.4. This allowed him to reduce the analysis
1.4. THE HOMOCLINIC TANGLE 19
Figure 1.5: The phase portrait of the damped pendulum is similar to the undamped
one except that the origin has changed to a sink and all solutions, except the saddle
and the two rotations that connect to the stable manifolds of the saddle, spiral into
the sink.
20 CHAPTER 1. INTRODUCTION
Figure 1.6: The plane perpendicular to the periodic orbit is called a transversal.
The Poincaré map is the return map to the transversal. Thus the Poincaré map
of the point q is the point p. The periodic orbit itself gives a fixed point of the
Poincaré map.
1.4. THE HOMOCLINIC TANGLE 21
Exercise 1.1
Perform the qualitative analysis of the Duffing’s equation,
ẍ − x + x3 = 0,
4. Use the phase portrait and (1) - (3) to identify four different types of solu-
tions of the Duffing’s equation and describe their qualitative behaviour.
5. Find the energy of the Duffing’s equation and use it to compute the homo-
clinic orbits, compare Appendix A.
22 CHAPTER 1. INTRODUCTION
Figure 1.7: The phase space of the Poincaré map of the damped and driven non-
linear pendulum.
1.4. THE HOMOCLINIC TANGLE 23
ẍ = δx − x + x3 = 0,
and describe how this changes the analysis (1) - (3) above when δ is small.
ẍ = δx − x + x3 = ε cos(ωt),
and describe what it’s Poincaré map looks like for small ε.
24 CHAPTER 1. INTRODUCTION
Chapter 2
25
26 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
then there exists a unique solution of (2.1) for the time interval |t − t0 | ≤
d
max (a, M ), where
M = max k f (x,t)k.
G
Proof: We write the ODE and the initial condition (2.1) as the integral equation
Z t
x(t) = x0 + f (x(s), s)ds. (2.2)
t0
It is easy to see that (2.1) and (2.2) are actually equivalent. Picard’s idea was to
iterate this equation and define the sequence
Z t
x1 (t) = x0 + f (x0 , s)ds
t0
Z t
x2 (t) = x0 + f (x1 (s), s)ds
t0
..
. Z t
xn+1 (t) = x0 + f (xn (s), s)ds
t0
..
.
If the sequence of iterates {xn (t)} converges it will converge to the solution of
(2.2). The first thing we must check is that the iterates actually lie in D, see Figure
2.1,
Z t
kx1 (t) − x0 k = k f (x0 , s)dsk
t0
Z t
≤ k f (x0 , s)kds
t0
≤ M|t − t0 | ≤ d
d
for |t − t0 | ≤ min (a, M ). This was the reason for the choice of the length of the
time interval, as this minimum. Now the computation is the same for all the
iterates so the whole sequence lies in D.
Next we use induction to show that successive iterates satisfy the inequality
MLn−1 |t − t0 |n
kxn (t) − xn−1 (t)k ≤ .
n!
2.1. THE PICARD EXISTENCE THEOREM 27
We assume
MLn−2 |t − t0 |n−1
kxn−1 (t) − xn−2 (t)k ≤
(n − 1)!
and consider the difference of
Z t
xn (t) = x0 + f (xn−1 (s), s)ds
t0
and Z t
xn−1 (t) = x0 + f (xn−2 (s), s)ds,
t0
Z t
xn (t) − xn−1 (t) = [ f (xn−1 (s), s) − f (xn−2 (s), s)] ds
t0
so
Z t
kxn (t) − xn−1 (t)k ≤ k f (xn−1 (s), s) − f (xn−2 (s), s)kds
t0
Z t
≤ L kxn−1 (s) − xn−2 (s)kds
t0
|s − t0 |n−2 |s − t0 |n
Z t
n−1
≤ ML ds = MLn−1 t
t0
t0 (n − 1)! n!
|t − t0 |n
= MLn−1
n!
This completes the induction.
Now the space of continous funtions on the interval I is a complete metric
space with the metric
kx(t)ksup = max |x(t)|
I
and to show that the sequence {x(t)} converges we just have to show that it is
Cauchy, i.e. given ε > 0, ∃N such that
M ∞
L j |t − t0 | j M L|t−t0 |
≤ ∑ j! = e − 1 .
L j=1 L
ẋ = x1/2
x(0) = 0.
x ≡ 0 is a solution but so is
(t − c)2 /4, t ≥ c > 0
x(t) =
0, t < c.
We have infinitely many solutions one for each value of c, see Figure 2.1.
30 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
Figure 2.2: The top figure show the graphs of two of the many solutions that
satisfy the ODE in Example 2.1. The bottom figure show the solution in Example
2.2, that is blowing up in a finite time.
2.1. THE PICARD EXISTENCE THEOREM 31
ẋ = x2
x(0) = x0 .
The solution is
x0
x(t) = .
(1 − x0t)
It exists for t < 1/x0 , but
lim x(t) = +∞,
t→1/x0
Exercise 2.1
has infinitely many continuous solutions in the region |x| ≤ 1. Show also
that there are only two solutions that have a continuous second derivative ẍ.
ẋ = 1 + x2
x(0) = x0
Theorem 2.2 Let u(t) and v(t) be non-negative continuous functions on [α, β],
C ≥ 0 a constant, and suppose
Z t
v(t) ≤ C + v(s)u(s)ds, for α ≤ t ≤ β. (2.3)
α
32 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
Then
t
Z
v(t) ≤ C exp u(s)ds , for α ≤ t ≤ β (2.4)
α
and if C = 0, v(t) ≡ 0.
Proof: We let
Z t
C+ v(s)u(s)ds = V (t).
α
Then
v(t) ≤ V (t)
by the hypothesis and we assume in addition that
then
V 0 (t) = v(t)u(t) ≤ V (t)u(t).
We can solve this differential inequality, because
The solution is
t t
Z Z
V (t) ≤ V (0) exp u(s)ds = C exp u(s)ds .
α α
and
C = 0 ⇒ v(t) ≡ 0 for (t − α) small.
But then v(t) ≡ 0, for all t ∈ [α, β] because we can make the same argument with
any α1 ∈ [α, β] instead of α. QED
Corollary 2.1 Let y(t) and x(t) be two solutions of the ODE (2.1), with a Lips-
chitz continuous vector field f and assume that the solution exists for |t −t0 | < T0 .
Then for ε > 0 and T < T0 , there exists a δ > 0 such that
for |t − t0 | ≤ T .
Now
ky0 − x0 k < εe−Kt = δ
implies that
ky(t) − x(t)k < ε, for |t − t0 | < T.
QED
The next application of the Grönwall’s inequality is to show that if the deriva-
tive of the vector field is globally bounded, then the solution is bounded by an
exponential function. First we must discuss:
If the vector field is only assumed to be continuous the solution of the Initial
Value Problem (2.1), but it may not be unique. The proof requires the classical
Arzela-Ascoli Theorem: A uniformly bounded and equicontinuous sequence of
functions on a compact set G ⊂ Rn+1 has a uniformly convergent subsequence.
ẋ = f (x,t), x(t0 ) = x.
Example 2.3
The motion of a spring with a stiffening (nonlinear) restoring force is described
by the equation
ẍ = −x − x3
We move the two terms on the right hand side left and multiply by ẋ. Then inte-
gration in t yields the energy
ẋ2 x2 x4
+ + = constant = K
2 2 4
This implies that
|x| ≤ 2K 1/2 , and |ẋ| ≤ (2K)1/2
for all time, so the solution exist globally, see Corollary 2.3 below.
ẍ + sin(x) = 0
Exercise 2.3 What can you say about the interval of existence of the solution to
ẍ + x2 = c
where c is a constant?
Global solutions are obtained by piecing together local solutions. In fact it
follows from the Local Existence Theorem 2.1 that
1. Every point (to , xo ) lies on some solution (t, x(t)).
2. If (to , xo ) lies on two solutions (t, x(t)) and (t, y(t)), then these solutions
must be the same in an interval |t − to | < a, around to .
The second statement above is strengthen considerably by the follwing Lemma
Lemma 2.1 If two solutions of the same ODE agree at a point then they must be
the same in their whole interval of existence.
Proof: Suppose that the solution (t, x(t)) is defined on the interval [a, b) and the
other solution (t, y(t)) is defined on the interval [a, c) with b < c. Let
d = in f {t|x(t) 6= y(t)}
and consider the interval [a, d]. x(t) and y(t) both exist on [a, d) and since they are
continuous, their limits at d must be equal
This means that d > b and consequently b = c because otherwise x(t) exists be-
yond b contradicting the statement above. QED
We define the solution in the large (t, z(t)) to be the union of the local solutions
(t, x(t)). It exists on the union of the local intervals and is uniquely defined by any
initial point (to , xo ) by above discussion.
Next we prove in two steps that if solution only exist for a finite time interval
then they must blow-up.
Theorem 2.5 Let D̄ be a closed and bounded subset of a region U where solution
(t, x(t)) is defined. If the solution is only defined on the time interval [a, b) with
b < ∞, then (t, x(t)) must leave D̄ for all t sufficiently close to b.
2.2. GLOBAL SOLUTIONS 37
Figure 2.3: The compact set D̄ has the distance γ to the compliment of the region
U.
38 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
Proof: We assume there exists a solution (t, x(t)) defined for 0 ≤ |t − t0 | < β and
passing through some point (t0 , x0 ) of D̄. We have to prove that there exists an
α such that the solution lies outside D̄, for β − α < t − t0 < β. Now consider the
region Rn+1 \U, see Figure 2.2. Let
be the distance between D̄ and Rn+1 \U, and consider the rectangles
Then by the Picard Existence Theorem 2.1 (t, x(t)) exists in the interval |t − t0 | <
α. Now if β − α < t1 < β and (t1 , x(t1 )) ∈ D∗ then the solution would exist beyond
t − t0 = β which is a contradiction. QED
Corollary 2.3 If the solution x(t) of the IVP only exists for time less than b < ∞,
then
lim kx(t)k = ∞
t→b
Then (t, x(t)) must leave D̄ for t approaching b, but since it cannot leave the side
t = b of D, it must leave the sides kxk = d. However, d is arbitrary. QED
2.3. LYAPUNOV STABILITY 39
Definition 2.2 A solution of the first order system (2.6) is Lyapunov stable if for
ε > 0, there exists δ > 0, such that
1. V is non-negative, V (x) ≥ 0
3. V is non-increasing
dV
≤0
dt
Theorem 2.6 Suppose the vector field f (x) is continuous in a neighborhood of
a stationary solution x̄ and that there exists a Lyapunov function for (2.6) in this
neighborhood. Then x̄ is stable. Morover, if
dV
<0
dt
then x̄ is asymptotically stable.
40 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
We define exit points and prove a technical lemma before proving the theorem.
Lemma 2.2 Let f be continuous on an open subset U and D ⊂ U open such that
∂D ∩ U is either empty or consists of points which are not exit points. Then the
solution of (2.6), with (t0 , x0 ) ∈ D, stays in D as long as it exists.
Proof: Suppose that the solution (t, x(t)) leaves D at some time t1 , then
(t1 , x(t1 )) ∈ ∂D ∩U
Proof: Choose ε > 0 and let γ < minkx−x0 k=ε V (x). We can then choose δ such
that the ball kx − x0 k ≤ δ lies inside the region V (x) = γ, see Figure 2.3. We define
the sets
U = {x| kx − x̄k < ε}, D = {x| V (x) < γ}, ∂D = {x| V (x) = γ}
v̇ = V̇ ≤ 0
Example 2.4
2.3. LYAPUNOV STABILITY 41
Figure 2.4: The set defined by V (x) = γ lies outside the δ ball and inside the ε ball.
42 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
ẋ = σ(y − x)
ẏ = ρx − y − xz
ż = xy − βz
1
V (x, y, z) = (x2 + σ(y2 + z2 ))
2
is a Lyapunov function for the Lorenz equations. First V ≥ 0 is positive definite
since σ > 0, V (x, y, z) = 0 implies (x, y, z) = (0, 0, 0) and the origin is a stationary
solution of the Lorenz equations. This verifies the first two conditions of Defini-
tion 2.2. By use of the Lorenz equations
dV
= xẋ + σ(yẏ + zż)
dt
= σ(xy − x2 ) + σ(ρxy − y2 − xyz) + σ(−βz2 + xyz)
Thus
Definition 2.5 The ODE, ẋ = f (x), has an absorbing set D; if for every bounded
set U ⊂ Rn , there exists a time T (U) such that x(0) = xo ∈ U and t ≥ T (U) implies
that x(t) ∈ D.
This means that all orbits are eventually absorbed by D see Figure 2.4.
Example 2.5
Consider and ODE
ẋ + δx = f (x,t),
where k f (x,t)k ≤ K. We can integrate this equation to get
d δt
e x = eδt f (x,t)
dt
or Z t
−δt
x(t) = xo e + e−δ(t−s) f (x(s), s)ds
0
and
K
kx(t)k ≤ kxo ke−δt + (1 − e−δt ). (2.7)
δ
The last inequality implies that
K
kx(t)k ≤ +κ
δ
where κ is an arbitrarily small number, defines an absorbing set for the ODE.
Namely, if we solve the inequality (2.7) for t, we get
K K
kxo ke−δt + (1 − e−δt ) ≤ + κ
δ δ
and
K −δt
(kxo k − )e ≤ κ
δ
so
κ
−δt ≤ log[ ]
(kxo k − Kδ )
and
1 κ
t ≥ − log[ ].
δ (kxo k − Kδ )
44 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
Figure 2.5: Every orbit starting in U is eventually absorbed by the absorbing set
D.
2.4. ABSORBING SETS, OMEGA-LIMIT SETS AND ATTRACTORS 45
This means that if the initial data lies in a bounded set defined by kxo k ≤ M, then
x(t) ∈ D for
1 κ
t ≥ T (M) = − log[ ].
δ (kxo k − Kδ )
Stationary solutions are the simplest structure that one can encounter in phase
space the ω and α limit sets are more general structures. Stationary solutions
are clearly invariant under the flow and the union of orbits tending towards them
in positive time is called their basin of attraction. We will now generalize these
notions to more complicated attracting sets.
Definition 2.6 The ω limit set of a solution to an ODE, x(t), x(0) = xo consists
of all points y such that there exists a sequence tn → ∞ as n → ∞ and
lim x(tn ) = y.
n→∞
The α limit set of x(t) consists of all points y such that there exists a sequence
tn → −∞ as n → ∞ and
lim x(tn ) = y.
n→∞
The ω limit set consists of all points that points sampled from the solutions
converge to in positive time, whereas the α limit set consists of the points that
points sampled from the solution converge to in negative time.
Theorem 2.7 The α and ω sets are closed sets, if the solution x(t) is also compact
then these sets are compact, non-empty and connected sets.
Proof: We first show that the ω limit set is closed. Suppose that
lim yn = y
n→∞
where yn ∈ ω{x(t)}. Then by definition of ω there exist sequences {tmn } such that
lim x(tmn ) → yn .
n→∞
We can choose a (diagonal) sequence {tnn } from the sequences {tmn } such that
lim x(tnn ) → y.
n→∞
Thus y ∈ ω{x(t)}.
46 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
and thus
lim x(tk ) = z,
k→∞
so z ∈ ω{x(t)}. The distance
d
d(X, z) ≥ d(X,Y ) − d(Y, z) = ,
2
so z is neither in X nor Y , but this contradicts the decomposition
ω = X ∪Y.
The proof for the α limit set is similar. QED
2.4. ABSORBING SETS, OMEGA-LIMIT SETS AND ATTRACTORS 47
Exercise 2.4 Show that ω(xo ) and α(xo ) are both invariant.
Theorem 2.8 If an ODE, ẋ = f (x), has an absorbing set D, then the ω-limit set
of D
ω(D) = ∩n ∪m≥n T m D
is a global attractor.
Proof: We first observe that ω(D) is compact and non-empty because it is a
nested intersection of closed and bounded sets. Then we show that it is invariant.
Let Kn = ∪m≥n T m D and K = ω(D). Then K is positively invariant because if
y ∈ K, there exists a sequence zn = T n yn → y and T m zn = T m+n yn ∈ K m+n , thus
T m y = limn→∞ T m+n yn ∈ K.
K is also negatively invariant. If T n yn → y ∈ K then {T n−m ym } has a convergent
subsequence, by the compactness of K, and if z = limn→∞ T n−m ym , then
y = T m z ∈ T m K.
This shows that K ⊂ T m K. Now K attracts a neighborhood, because if it does not,
then there exists a sequence yn in a neighborhood of K such that
kT n yn − Kk ≥ ε > 0.
But {T n yn } has a convergent subsequence
y = limm→∞ T m ym ∈ K,
which is a contradiction. Moreover, K attracts D which attracts Rn so K attracts
Rn . QED
48 CHAPTER 2. EXISTENCE, UNIQUENESS AND INVARIANCE
Theorem 2.9 If an ODE, ẋ = f (x), has a Lyapunov function; then its global at-
tractor is connected and can be expressed as
A = W u (S),
A = ∪N u
x̄ j ∈SW (x̄ j ).
Lemma 2.3 If the ODE, ẋ = f (x), x(0) = xo , has a Lyapunov function and xo ∈
Rn , then ω(xo ) ∈ S and if x(t), is a compact orbit then α(xo ) ∈ S, also.
x(tn + t) → y,
V (x(tn + t)) → c.
Then
v =V −c
is also a Lyapunov function and v(y) = 0. This means that y = x̄ ∈ S. QED
QED
The Picard Theorem 2.1 is an example (the first in history) of a fixed point or
contraction mapping theorem.
T x = x.
Proof: We first prove the uniqueness. Suppose x and y are both fixed points
T x = x, Ty = y. Then
θn
d(T m x, T n x) ≤ d(T x, x), m > n
(1 − θ)
Now sending n → ∞ we see that the sequence is Cauchy and since the metric space
is complete, there exists an x in the space such that
lim T n x = x.
n→∞
QED
Chapter 3
(y1 , y2 , . . . , ym ) = f (x1 , x2 , . . . , xn )
1. ∪ j∈JU j = M
2. ϕ j o ϕ−1 k
k is C , for j, k ∈ J
51
52 CHAPTER 3. THE GEOMETRY OF FLOWS
Example 3.1
The n-sphere is a n-dimensional manifold.
n
Sn = {x ∈ Rn+1 | ∑ x2j + (xn+1 − 1)2 = 1}
j=1
or the sphere with the two poles removed. The corresponding coordinate functions
are the stereographic projections from each punctured sphere ϕ1 = p1 , ϕ2 = p2 ,
see Figure 3.1. In physics the manifold M is the phase space of the physical system
and the motion on M is described by the ODE.
2. gt gs = gt+s
3. g0 = identity
These properties imply that gt−1 = g−t . For t = T fixed, gT is nothing but the
time-T map, see Definition 2.7.
Definition 3.7 The one parameter group along with its manifold (M, {gt }) is
called the (phase) flow.
3.1. VECTOR FIELDS AND FLOWS 53
Figure 3.1: The two sphere with the two covering sets U1 and U2 and correspond-
ing steriographic projection p1 and p2 to R2 .
54 CHAPTER 3. THE GEOMETRY OF FLOWS
Exercise 3.1
1. Find the phase space of the harmonic oscillator
ẍ + ω2 x = 0
and the one parameter group of diffeomorphisms gt . Also find two different
type of orbits. Are these orbits phase curves?
2. Consider the nonlinear pendulum
ẍ + sin(x) = 0
Describe four different types of phase curves for the nonlinear pendulum.
We differentiate the orbit with respect to time to get the phase velocity
φ̇t (x) = v(x)
Since the orbit φt (x) = x(t) is just the solution the ODE, ẋ = f (x), with initial
condition x(0) = x, we conclude that the phase velocity is just the vector field
v(x) = f (x)
This is where the name vector field originates, v is the vector field on the phase
space M that determines the flow.
3.1. VECTOR FIELDS AND FLOWS 55
Example 3.2
1. ẋ = kx
2. ẋ = sin(x)
1. The phase space is M = R, the one-parameter group of diffeomorphism is,
gt = ekt
and the orbit is
φt (x) = ekt x
x̄ = 0 is the single stationary solution and singular point of the vector field.
The flow is away from 0, so 0 is a unstable stationary solution, see Figure
3.1.
2. The phase space is M = R, but gt is not a simple function, see the Exercise
3.2. However, gt and φt (x) exist and can be worked out. The stationary
solutions and singular point of the vector field are x = nπ, n ∈ Z. The even
multiples of π are unstable and the odd multiples are stable, see Figure 3.1.
Exercise 3.2
1. Find the second semi-group in Example 3.2
2. Find gt and φt (x) for Example 2.
3. Show that a one parameter group of transformation may not always exist or
only exists for a finite time.
56 CHAPTER 3. THE GEOMETRY OF FLOWS
Figure 3.2: The direction vector field tangent to the integral curve of the orbit.
3.1. VECTOR FIELDS AND FLOWS 57
Figure 3.3: The extended phase space and the phase space of the ODEs 1 and 2,
in Example 3.2, respectively.
58 CHAPTER 3. THE GEOMETRY OF FLOWS
v : T Mx → Rn
is an isomorphism.
Proof: We have to show that if two tangent vectors v1 = v2 are the same then the
corresponding orbits are tangent. Since
v1 − v2 = 0
Dx y (φ̇t1 − φ̇t2 ) = 0.
However, since Dx y is invertible it follows that
φ̇t1 − φ̇t2 = 0.
Dx g : TUx → TVy
g : W → g(W )
F(x, y) = y − g(x)
such that
F(x0 , y0 ) = y0 − g(x0 ) = 0.
60 CHAPTER 3. THE GEOMETRY OF FLOWS
Then
detDx F = −detDx g 6= 0,
at x = x0 , and by the implicit function theorem there exists a small neighborhood
E ⊂ V of y0 and a function
h : E → h(E) = W
such that
F(h(y), y) = 0
for y ∈ E. Moreover, h is unique and differentiable as often as g. Clearly
h = g−1 ∈ C1 ,
ẋ = f (x, θ)
θ̇ = 1
Thus our definition of the tangent vector and tangent space hold in the extended
phase space M × R with coordinates (x, θ). However, frequently this is not the
most desirable result, we would like objects such as the semi-group to be defined
for the phase space M itself and this is not always possible to do. But the addition
of θ as above can be a very useful tool and it is just what we need for the next
theorem. It is called the Rectification Theorem and it says that in a neighborhood
of a nonsingular point any flow can be mapped onto straight linear flow.
ẋ = f (x, θ)
θ̇ = 1
3.3. FLOW EQUIVALANCE 61
ẏ = 0
ż = 1
Proof: We define a map h from the line (y, z) to the orbit (φt (x), θ). If we can
show that h is a diffeomorphism, then g = h−1 is the desired map. Define
Moreover
detD(y,z) h = 1 6= 0
and by the inverse function theorem 3.1 h is a diffeomorphism. If v = f is Ck−1 ,
then h is Ck and g is also Ck by the inverse function theorem. QED
Exercise 3.3 Prove the Picard Theorem 2.1 (away from a singular point) using
the Rectification Theorem.
62 CHAPTER 3. THE GEOMETRY OF FLOWS
Figure 3.4: The two vector fields in the Rectification Theorem and the map be-
tween them.
3.3. FLOW EQUIVALANCE 63
Lemma 3.3 Two linear systems are linearly conjugate if and only if they have the
same eigenvalues with the same algebraic and geometric multiplicity.
Proof: Consider the systems
ẋ = Ax and ẏ = By
they are linearly conjugate if
φt2 (h(x)) = hφt1 (x)
where h is an n × n invertible matrix. Now
φ̇t2 = hφ̇t1 = hAφt1
= hAh−1 hφt1 = hAh−1 φt2
Comparing
φ̇t2 = hAh−1 φt2 and ẏ = By
we conclude that A and B are similar matricies
B = hAh−1
Conversely, if A and B are similar matricies then the equations
φ̇t2 = hAh−1 φt2 and ẏ = By
can both be solved with the same initial data to give the solution
φt2 (hx) = hφt1 (x)
by the uniqueness of the solution to the initial value problem. This shows that the
two orbits are linearly conjugate. QED
64 CHAPTER 3. THE GEOMETRY OF FLOWS
Theorem 3.3 Two linear systems are differentiably conjugate if and only if they
are linearly conjugate.
ẋ = Ax and ẏ = By
If the orbits are differentiably conjugate by Lemma 3.3 is suffices to show that A
and B are similar matricies. Suppose that
Similarly,
φt2 (x) = y + Byt + O(t 2 )
We substitute these expansions into Equation 3.2 and differentiate with respect to
the initial data x. This can be done because g is a diffeormophism and the orbits
depend continuously on their initial data. The differentiation of
gives
(I + Bt + O(t 2 ))g0 = g0 (I + At + O(t 2 ))
where g0 is the Jacobian of g. This shows that
Bg0 = g0 A
and the matricies are similar. The converse is immediate because a linear conju-
gacy is a C∞ diffeomorphism. QED
Remark 3.1 Theorem 3.3 shows that differentiable conjugacy is too strict an
equivalence relation at a singular point of a vector field. We must settle for topo-
logical conjugacy.
We will show below that generically a nonlinear flow is topologically con-
jugate to its linearization. The topological conjugacy classes are given by the
invariant manifold theorems. Namely, the stable and unstable manifold theorem
and the center manifold theorem.
Chapter 4
Invariant Manifolds
2. x(t0 ) ∈ W s =⇒ x(t) ∈ W s , ∀ t ≥ t0
65
66 CHAPTER 4. INVARIANT MANIFOLDS
Eu
u
W
s
W
Es
2. x(t0 ) ∈ W u =⇒ x(t) ∈ W u , ∀ t ≤ t0
3. x(t0 ) 6= W u =⇒ ∃ δ > 0, t1 ≤ t0 , kx(t)k > δ, ∀ t ≤ t1
Proof: By the statement of the theorem we both have to prove the existence of the
functions hs and hu whose graphs are the manifolds W s and W u and the existence
of the solutions to the ODE xs (t) ∈ W s and xu (t) ∈ W u .
If Es is the linear stable manifold of the linearized system ẏ = Ay then A leaves
Es invariant, AEs ⊂ Es , and if we define the exponential of a matrix to be the power
series
A2t 2 Ant n
etA = I + At + +···+ +··· ,
2 n
then the exponential also leaves Es invariant, etA Es ⊂ Es . Moreover, we have an
estimate
ketA ys k ≤ Cs e−ρt kys k, t ≥ 0 (4.1)
and the same remarks apply to Eu so etA Eu ⊂ Eu and
ketA yu k ≤ Cu eσt kyu k, t ≤ 0 (4.2)
Cs , and Cu are constants. −ρ and σ are respectively the smallest (in absolute value)
negative and positive real parts of the eigenvalues of A.
67
Since Rn is a direct sum Rn = Es ⊕Eu we can split x = xs +xu and write an integral
equation for xs and xu separately
Z t
(t−t0 )A
xs (t) = e xs (t0 ) + e(t−τ)A g(x(τ))s dτ
t0
and Z t
(t−t0 )A
xu (t) = e xu (t0 ) + e(t−τ)A g(x(τ))u dτ.
t0
We have used here that both Es and Eu are invariant under the flow. Next we
look for a solution that lies in W s and therefore remains in a neighborhood of the
origin for all t ≥ 0. If we consider the second equation and notice that xu picks up
the positive eigenvalues of A, we realize that the only way we are going to get a
solution that stays in a neighborhood of the origin is to send t0 → ∞. Then
0
lim ke−t0 A xu (t0 )k ≤ lim Cu0 e−σ t0 = 0,
t0 →∞ t0 →∞
assuming xu (t) stays bounded, where σ0 is now the smallest positive real part of
the eigenvalues of A. This gives
Z t
xu (t) = e(t−τ)A g(x(τ))u dτ
∞
Now let xs (0) = xs0 ∈ Es and recall some spaces that we already encountered
in the proof of Picard’s theorem. First let
C0 ([0, ∞); Rn )
be the space of continuous functions from the half-line R+ = [0, ∞) into Rn . Each
point in C0 ([0, ∞); Rn ) is a n-vector valued function x(t) ∈ Rn , for t fixed. How-
ever, C0 ([0, ∞); Rn ) is not a Banach space under the sup norm. It is a Fréchet
68 CHAPTER 4. INVARIANT MANIFOLDS
space or a complete metric space under a sequence of quasi-norms but this is not
what we need. We have to introduce a weight eµt where µ = min(ρ, σ) to make
Cµ0 ([0, ∞); Rn ) into a complete normed linear space, or in other words a Banach
space, under the norm
kx(t)kµ = sup eµt kx(t)k.
t∈[0,∞)
We have forced the function in Cµ0 ([0, ∞); Rn ) to decay exponentially in t by in-
serting the weight eµt and this makes the space a Banach space. We will restrict
the space further by making the functions lie in a small ball in Rn and define the
space on which we will prove the contraction to be
0
Cµ,δ ([0, ∞); Rn ) = {x(t) ∈ Cµ0 | kx(t)k ≤ δ}.
This restriction partially destroys the Banach space structure because the sum of
0 in no longer lies in this ball, x + y ∈ C0 , so it is not a linear
two vectors x, y ∈ Cµ,δ µ,δ
space. However, it is easy to check that this space is a complete metric space and
this is all we need for an application of the Contraction Mapping Principle 2.10.
Now we use the integral equation (4.3) to define a map
Z t Z ∞
(t−τ)A
tA
F (x(t)) = e xs (0) + e g(x(τ))s dτ − e(t−τ)A g(x(τ))u dτ. (4.4)
0 t
(4.3) by eµt and apply the triangle inequality and the two inequalities (4.1) and
(4.2) above, and the hypothesis on g, in the first step; since lim|x|→0 kg(x)k
kxk = 0, so
that there exists ε > 0 such that kDg(x)k ≤ εkxk,
Z t
−(ρ−µ)t
µt
e kx(t)k ≤ Cs e kxs (0)k +Cs ε e−(ρ−µ)(t−τ) eµτ kx(τ)kdτ
Z ∞ 0
≤ e−(ρ−µ)t kx1,s
0 0
− x2,s k
Cs Cu
+ ε +ε sup eµt kx1 (t) − x2 (t)k.
(ρ − µ) (σ + µ) [0,∞)
We have shown that the integral equation (4.3) has a unique solution x(t), but
this gives us a function hs : Es → Eu . Namely, x(t) = (xs , xu )(t) where
Z t
(t−t0 )A
xs (t) = e xs (t0 ) + e(t−τ) gs (x(τ))dτ
t0
and Z ∞
xu (t) = − e(t−τ)A gu (x(τ))dτ.
t
70 CHAPTER 4. INVARIANT MANIFOLDS
This is the desired function. The dependance of hs on xs enters throught the de-
pendance on x(τ) that depends on the initial data xs (t0 ) = xs .
We will now prove that the function hs has the properties (1) - (3). To prove
(1) we consider the integral equations above that xs (t) satisfies,
Z t
xs (t) = etA xs (0) + e(t−τ)A g(x(τ))s dτ,
0
where we have chosen t0 = 0 for convenience. We get using the triangle inequality
for the norm and taking the limit as t → ∞
Z t
−ρt −µt
lim kxs (t)k ≤ lim Cs e kxs (0)k + lim Cs εe e−(ρ−µ)(t−τ) eµτ kx(τ)kdτ
t→∞ t→∞ t→∞ 0
Cs ε
≤ Cs lim e−ρt kxs (0)k + lim e−µt kx(t)kµ = 0.
t→∞ (ρ − µ) t→∞
This integral converges uniformly, since Dxs x is bounded by the Continous De-
pendence on Parameters Theorem and
by the estimate
Z ∞
−µt
s
kDxs h (0)k = lim Dxs xu (t) ≤ lim Cu e e(σ+µ)(t−τ) eµτ kx(τ)kkDxs x(τ)kdτ
t→∞ t→∞ t
C
≤ lim e−µt kx(t)kµ = 0.
(σ + µ) t→∞
To prove (2) we consider the integral equation (4.3). If x(t0 ) ∈ W s then
Z ∞
xu (t0 ) = − e(t−τ)A g(x(τ))u dτ,
t0
and xs (t0 ) = xs . Solving the equations for xu (t) and xs (t) with this initial data, we
get Z t
xs (t) = e(t−t0 )A xs + e(t−τ)A g(x(τ))s dτ (4.5)
t0
Z ∞
xu (t) = − e(t−τ)A g(x(τ))u dτ. (4.6)
t
However, adding the two equations (4.5) and (4.6) and setting t0 = 0, gives the
equation (4.3) that characterizes W s so x(t) ∈ W s for t ≥ 0.
The property (3) is proven by an estimate. If we write the original ODE in
integral form then
Z t Z t
(t−t0 )A (t−τ)A (t−t0 )A
x(t) = e xu (t0 )+ e g(x(τ))u dτ+e xs (t0 )+ e(t−τ)A g(x(τ))s dτ.
t0 t0
We let σ0 be the smallest positive real part of the eigenvalues of A, σ is the largest
positive real part of the eigenvalues of A as above. Now by the triangle inequality
and the estimates (4.1) and (4.2)
Z t
σ0 (t−t0 ) −µt
kx(t)k ≥ Cu e kxu (t0 )k −Cu εe e(σ+µ)(t−τ) eµτ kx(τ)k)dτ
t0
Z t
−ρ(t−t0 ) −µt
− Cs e kxs (t0 )k −Cs εe e−(ρ−µ)(t−τ) eµτ kx(τ))kdτ
t0
0 ε
≥ Cu (eσ (t−t0 ) kxu (t0 )k − e−µt kx(t)kµ ) −Cs (e−ρ(t−t0 ) kxs (t0 )k
(σ + µ)
ε
+ e−µt kx(t)kµ ) ≥ δ,
(ρ − µ)
for t − t0 sufficiently large. This shows that if xu (t0 ) 6= 0 so x(t0 ) 6∈ W s then x(t)
must leave a neighborhood of the origin at some time t1 ≥ t0 and stay out for t ≥ t1 .
QED
72 CHAPTER 4. INVARIANT MANIFOLDS
How do we compute the stable and the unstable manifolds? We derive an equation
for hs (and hu ) and approximate these function by quadradic and higher Taylor
polynomials. The equations can be written as
ẋ = Ax + f (x, y), x ∈ E s
ẏ = By + g(x, y), y ∈ E u .
Dx hs ẋ = By + g(x, hs )
or
Dh(Ax + f (x, h)) = Bh(x) + g(x, h)
where we dropped the superscript and all functions are functions of x only. This
is the equation satisfied by hs (x), similarly hu (y) satisfies the equation
Example 4.1 2
d x1 −1 0 x1 −x2
= +
dt x2 0 1 x2 x12
ẋ = −x − y2 A = −1
ẏ = y + x2 B=1
or in our case
h0 (−x − h2 ) = h + x2 .
We look for a solution that is a power series in x
or
−2ax2 − 3bx3 + · · · = (a + 1)x2 + bx3 + · · · .
Equating coefficients gives
−2a = a + 1, a = −1/3
−3b = b, b = 0.
Thus
−x2
hs (x) = + O(x4 ).
3
Exercise 4.1
1. Find the linear and nonlinear stable and unstable manifolds of the hyper-
bolic stationary solutions of the Duffing equations,
ẍ ± (x − x3 ) = 0,
2. Find the stable and unstable manifolds of the hyperbolic stationary solution
at the origin for the system
3
d x 1 0 x −y
= +
dt y 0 −1 y x3
74 CHAPTER 4. INVARIANT MANIFOLDS
Chapter 5
Chaotic Dynamics
is a Ck diffeomorphism if f : Rn −→ Rn is a Ck diffeomorphism.
The orbit of the map {xm } ⊂ Rn is now a sequence generated by composing
the function f with itself. We will use the shorthand
f m (x) = f ◦ f ◦ · · · ◦ f (m times).
Maps have their local existence theory analogous to the ODE theory in Chapter I
and a linear theory analogous to the theory of linear ODE’s. We will not review
this theory here. A nice account of the theory of linear ODE’s can be found in
Perko [17] Chapter 1.
We will use diffeomorphism as a tool to understand flows (one can also use
flows as a tool to understand diffeomorphisms) and in particular, we are interested
in the analog for maps of stationary solutions of ODE’s and their stability.
75
76 CHAPTER 5. CHAOTIC DYNAMICS
xm+1 = f (xm )
where
f (x) = 1 − µx2 ,
see Figure 5.1. This map has two fixed points where the graph of f meets the line
y = x and since one of these points has slope less than one, it is stable. The other
has slope greater than one and is unstable, see below.
Definition 5.3 An orbit {xm } of a map (5.1) is stable if for ε > 0, there exists
δ > 0, such that
lim ym = xm .
m→∞
Lemma 5.1 If |λ| < 1 for all λ ∈ σ(D f (x)) then x is (asymptotically) stable. If
|λ| > 1 for some λ ∈ σ(D f (x)) then x is unstable.
Here σ(D f (x)) denotes the spectrum of f linearized about the fixed point x.
Proof: By the Mean-Value Theorem
Z 1
xm − ym = D f (sxm−1 + (1 − s)ym−1 ) ds · (xm−1 − ym−1 ).
0
lim ym = x.
m→∞
78 CHAPTER 5. CHAOTIC DYNAMICS
If there exists an eigenvalue |λ| > 1 then we can pick x−y0 to be the corresponding
eigenvector. Then
Z 1
x − y1 = D f (sx + (1 − s)y0 ) ds · (x − y0 ),
0
kx − y1 k ≥ (|λ| − η)kx − y0 k,
kx − ym k ≥ (|λ| − η∗ )m kx − y0 k,
where η∗ is still small. This shows that {ym } diverges from x. QED
If x∗ is a periodic orbit of period m then we get the criteria,
Corollary 5.1 If |λ| < 1 for all λ ∈ σ(D f m (x∗ )) then x∗ is (asymptotically) stable.
If |λ| > 1 for some λ ∈ σ(D f m (x∗ )) then x∗ is unstable.
Invariant sets for maps are defined in the same way as those for flows in Chap-
ter 1. As for ODE’s we are interested in conjugacy classes of maps.
h◦g = f ◦h
where h is a homeomorphism.
Proof: Since f is a diffeomorphism and D f (x) > 0 for some x, it follows that
D f (x) > 0 for all x ∈ R. Consider the graph of f and its intersections with the line
y = x in Figure 5.1. The fixed points of the map xi = f (xi ) are clearly equal to the
stationary solutions of the ODE
ẋi = f (xi ) − xi = 0.
Moreover the orientation or the direction in which the points map or flow is the
same for the map and the flow in each subinterval [xi , xi+1 ]. This is because sign
5.1. MAPS AND DIFFEOMORPHISMS 79
Figure 5.2: The graph of the function f (x) and the line y = x.
80 CHAPTER 5. CHAOTIC DYNAMICS
[xm+1 − xm ] = sign [ f (xm ) − xm ] = sign ẋ. Now let x0 ∈ (xi , xi+1 ) be any point and
define the intervals
f : [Pm , Pm+1 ] −→ [Pm+1 , Pm+2 ]
where Pn = f n (x0 ) and
ϕ1 : [Qm , Qm+1 ] −→ [Qm+1 , Qm+2 ]
where Qm = ϕm (x0 ). Both f and ϕ1 move the intervals in the same direction.
We want to construct a homeomorphism taking the orbits of ϕ onto the orbits of
f . First we construct a homeomorphism taking [Q0 , Q1 ] onto [P0 , P1 ]. There are
many ways of doing this but we will pick the simplest and let
f (x0 ) − x0
h0 (y) = x0 + [y − y0 ].
ϕ1 (y0 ) − y0
Now clearly h(y0 ) = P0 and h(y1 ) = h(ϕ(y0 )) = f (x0 ) = P1 , so h[Q0 ] = P0 and
h[Q1 ] = P1 . Moreover, h is monotone (linear) and thus maps [Q0 , Q1 ] onto [P0 , P1 ]
in a one to one and invertible fashion. For the mth intervals we simply define
hm (y) = f m ◦ h0 ◦ ϕ−m (y).
In other words hm pulls [Qm , Qm+1 ] back to [Q0 , Q1 ] by ϕ−m , maps it to [P0 , P1 ] by
h0 and hm moves it forward to [Pm , Pm+1 ] by f m . The desired homeomorphism is
then
xi for y = xi
h(y) = hm (y) for y ∈ (Qm , Qm+1 )
xi+1 for y = xi+1 .
ẏ = D f (x)y, y = x − x,
ẋ = f (x)
xm+1 = f (xm ), m ∈ Z,
82 CHAPTER 5. CHAOTIC DYNAMICS
Figure 5.3: The positions of the eigenvalues and the corresponding motion of
the iterates of a one-dimensional map in phase space. The first two maps are
orientation preserving and the last two orientation reversing.
84 CHAPTER 5. CHAOTIC DYNAMICS
Exercise 5.1 Draw the eigenvalue configurations and map phase portrait for the
8 two-dimensional cases.
Heuristically, the stretching and compressing accounts for the hyperbolic part of
the map and the bending makes it nonlinear. Notice that the preimages of H0 and
H1 are vertical strips V0 and V1 and now we define the horseshoe map on S,
f (S) ∩ S = H0 ∪ H1
f (V j ) = H j , j = 0, 1.
The precise form of the map on SV0 ∪V1 is not crucial, it is linear on V j , j = 0, 1.
We take it to be as in Figure 2 and this becomes important when we embed the
map later in the sphere. The image of H j , j = 0, 1, in S, consists of four horizontal
strips H jk , j, k = 0, 1, see Figure 5.3. In general, the image of H n = f (H n−1 ) ∩
S, H 0 = H0 ∪ H1 consists of 2n horizontal strips which are contained in H0 and
H1 . The width of these strips is ≤ 1/5n −→ 0 as n −→ ∞. Thus the intersection of
all of these strips n∈Z+ H n is the product of the interval [0,1] and a Cantor set.
T
The inverse map similarly defines vertical strips, see Figure 5.3. f −1 is really
only defined on the horizontal strips H0 ∪ H1 , but we define
V 0 = V0 ∪V1 = f −1 (H0 ∪ H1 )
5.3. HORSESHOE MAPS AND SYMBOLIC DYNAMICS 85
Figure 5.4: The Horseshoe Map consists of stretching the square in the x direction,
compressing it in the y direction, bending the resulting rectangle into a horseshoe
and intersecting the horseshoe with the original square
86 CHAPTER 5. CHAOTIC DYNAMICS
Figure 5.5: The map of the first two horizontal strips consists of four horizontal
strips.
5.3. HORSESHOE MAPS AND SYMBOLIC DYNAMICS 87
product of a Cantor set with the y-interval [0, 1] and we define the Smale horseshoe
to be \ \ \
Λ= Hn ∩ Vn = f n (H0 ∪ H1 ).
n∈Z+ n∈Z+ n∈Z
The intersection of the first four vertical and horizontal strips in shown in Figure
5.3.
We now establish the relationship of the horseshoe map and symbolic dynam-
ics. Consider the space Σ of binfinite sequences of two symbols {0, 1}. σ ∈ Σ is a
sequence of zeroes and ones infinite in both directions
σ = (· · · , 1, 0, 1, 1, 1, 0, 0, 1, 0, · · · ).
Figure 5.6: The map of the first two vertical strips consists of four vertical strips.
5.3. HORSESHOE MAPS AND SYMBOLIC DYNAMICS 89
Figure 5.7: The horseshoe is the intersection of the vertical and horizontal strips.
90 CHAPTER 5. CHAOTIC DYNAMICS
Figure 5.8: The Smale Horseshoe lies with in the 16 squares that are labeled
according to which horizontal and vertical strips they came from.
5.3. HORSESHOE MAPS AND SYMBOLIC DYNAMICS 91
But these strips are nested H 0 ⊃ H 1 ⊃ · · · ⊃ H n and the forward symbol se-
quence
(σ0 , σ1 , σ2 , · · · )
restricts x to lie in a unique line Hσ0 σ1 ··· of the horizontal Cartesian product of
a line with a Cantor set. This is because a nested sequence of strips contains a
unique line. Moreover, we can repeat this labelling with the backward sequence
and the vertical strips Vσ−1 ···σ− j . Then the sequence segment
restricts x to lie in the rectangle Vσ−1 ···σ−( j+1) ∩ Hσ0 ···σ j and since these rectangles
are also nested and Vσ−1 ··· ∩Hσ0 ··· contains only one point, there is a 1:1 correspon-
dence between the points x ∈ Λ and the biinfinite sequences σ ∈ Σ. This proves
the following lemma.
Proof: Consider σ = h(x) and suppose x ∈ Vσ−1 ··· ∩Hσ0 ··· . Then f (x) ∈ Vσ−2 σ−3 ··· ∩
Hσ−1 σ0 σ1 ··· or
h ◦ f = α ◦ h.
QED
The dynamics on the sequence of two symbols can now be used to give a
complete description of the dynamics on the horseshoe.
92 CHAPTER 5. CHAOTIC DYNAMICS
Fact 5.1 There are infinitely many periodic orbits of all periods in Λ.
Just consider the sequences σ = (σ j ), σ j = 0 and σ = (σ j ), σ j = 1. These
correspond to fixed points of f on Λ, by the topological conjugacy, since ασ = σ
where α is the left shift on Σ. Then consider σ = (σ j ), σ j = 0, if j is even, σ j = 1
if j is odd. This is a periodic orbit of period 1, α2 σ = σ, and there is another one
namely σ j = 1, if j is even, σ j = 0 if j is odd. Thus Λ has exactly two orbits of
period 1 under f . Next we can make a periodic orbit out of a segment of length
three say 011, it is
σ = (· · · , 0, 1, 1, 0, 1, 1, 0, 1, 1, 0, 1, 1, · · · )
and there are 8 = 23 such because we have choice of two symbols for each σ j , j =
0, 1, 2. However, this counts the fixed points (but not the periodic orbits) above so
there are 6 = 8 − 2 periodic orbits of genuine period 2. It is clear that by taking a
segment of length 4 etc, we can construct periodic orbits of any period and count
them all.
A Metric
Now we define a metric on the space of sequences
∞
kσ1 − σ2 k = 2
∑ (σ1j − σ2j )2 /2| j| .
j=−∞
the positive direction. Then we take all segments of lenght three and string them
together. They will make the next segment of σ in the negative direction and so
on. Thus we get a sequence that contains arbitrarily large centrally located finite
segments of any sequence σ1 . Now let ε > 0 then there exist integers N and m
such that
kσ1 − αN σk = ∑ (σ1j − σ2j−N )2/2 j ≤ 1/2m−1 < ε,
| j|>m
because the centrally located, i.e. around the zeroth place, segments of σ1 and
αN σ, σ shifted left N times, of lenght 2m are identical.
Chaos
Fact 5.5 Λ contains chaotic orbits topologically conjugate to random flips of a
coin.
In other words: Λ contains points the orbits of which under horseshoe map are
topologically conjugate to a shift on a random sequence of two symbols.
We construct a sequence σ by flipping a unbiased coin, i.e. σ j = 0 for heads,
σ j = 1 for tails, j ∈ Z. Λ contains a sequence { f n (x)} n ∈ Z, which is topo-
logically conjugate to σ. This is the precise mathematical meaning of a chaotic
orbit.
Fact 5.7 The inset and outset of Λ in the square S are the Cartesian product of
the Cantor sets and lines,
\ \
in (Λ) ∩ S = V n , out (Λ) ∩ S = H n.
n∈Z+ n∈Z+ ∪{0}
Proof: Let x ∈ n∈Z+ V n Λ and consider the contraction of the horizontal strips
T
in Figure 5.3. The map f contracts the horizontal strips onto n∈Z+ ∪{0} H n and
T
shows that \ \
y = lim f m (x) ∈ Vn ∩ Hn = Λ
m→∞
n∈Z+ n∈Z+ ∪{0}
or y ∈ Λ. Similarly if
lim f −m (x) ∈ Λ.
\
x∈ H n Λ,
m→∞
n∈Z+∪{0}
QED
Exercise 5.3
1. The sixteen squares on Figure 5.3 that are the intersections of the four hori-
zontal and the four vertical strips contain each an initial point of a periodic
sequence of period three.
(a) Find the symbolic sequence for each of these periodic orbit and show
in which square it lies.
5.4. THE SMALE-BIRKHOFF HOMOCLINIC THEOREM 95
(b) How many of these are genuine orbits of period three? What are the
others?
(c) Describe how each initial point of a period three orbit moves around
some of the sixteen squares under the horseshoe map.
Lemma 5.7 The stable and unstable manifolds of the Poincaré map of (5.3) in-
tersect transversely if and only if the Melnikov function
Z ∞
M(to ) = f (xo (t − to )) ∧ g(xo (t − to ),t)dt (5.4)
−∞
has simple zeroes.
5.5. THE MELNIKOV METHOD 97
Figure 5.9: The unstable manifold must intersect the stable manifold transversely
in a point x0 6= x̄.
98 CHAPTER 5. CHAOTIC DYNAMICS
Figure 5.10: The homoclinic loop connect the stationary solution to itself W s =
W u.
5.6. TRANSIENT DYNAMICS 99
Exercise 5.4 Show that the Poincaré Map of the damped and driven Duffing’s
Equation
ẍ + δẋ − x + x3 = ε cos(t)
has a Smale Horseshoe in its phase space if
ε 4 cosh(π/2)
>
δ 321/2 π
Theorem 5.4 All points x of S n∈Z+ V n , where n∈Z+ V n is the vertical Carte-
T T
sian product of a Cantor set and the y-interval [0, 1], eventually approach the fixed
point in E,
lim F n (x) = p.
n→∞
Proof: Consider Figure 5.6. It shows that the regions A, B, D, and E are all
mapped into E in one iteration. All points in E are attracted to p. C is mapped into
A, but all of A gets mapped into E so C gets mapped into E in two iterations. Now
100 CHAPTER 5. CHAOTIC DYNAMICS
Figure 5.11: All points in a neighborhood, except the inset (stable manifold), of
the horseshoe, are eventually mapped to the sink.
5.6. TRANSIENT DYNAMICS 101
consider the two vertical strips V0 and V1 . The strips V 1 = V00 ∪V01 ∪V10 ∪V11 are
mapped onto V 0 = V0 ∪V1 in one iteration. This means that the points in V 0 V 1
must be mapped into B, C or D and therefore into E in at most three iterations, by
the above arguments. Similarly, V 1 V 2 is mapped into E in four iterations of F
etc. The remainder that does not get mapped into E is the intersection of all the
vertical strips n∈Z+ V n .
T
QED
102 CHAPTER 5. CHAOTIC DYNAMICS
Chapter 6
Center Manifolds
Recall from Chapters 1 and 4 the linear stable and unstable manifolds Es and Eu
of a hyperbolic stationary solution. Now we will also consider stationary solution
with pure imaginary eigenvalues and the linear center manifold Ec , spanned by the
eigenvectors of these pure imaginary or zero eigenvalues. The following theorem
give the existence of the nonlinear center manifold W c tangent to Ec .
ẋ = Ax + f (x, y), x ∈ Rl , y ∈ Rm ,
ẏ = By + g(x, y), l + m = n,
and assume that A has no pure imaginary (or zero) eigenvalues whereas B has
only pure imaginary (or zero) eigenvalues. Moreover, assume that f and g are
Ck (Rn ) functions so that
k f (x, y)k kg(x, y)k
lim = 0 = lim .
k(x,y)k→0 k(x, y)k k(x,y)k→0 k(x, y)k
h : Πc (U) −→ Es × Eu ,
hc (0) = 0, Dy hc (0) = 0.
103
104 CHAPTER 6. CENTER MANIFOLDS
where h0 = Dy h.
Proof: Substitute x = h(y) into the equation for x and use the equation for y to
eliminate ẏ. QED
ẋ = −x + x2 − y2
ẏ = εy − y3 + xy
ẋ = −x + x2 − y2
ẏ = εy − y3 + xy
ε̇ = 0
then there is a two dimensional center manifold. We let x = h(y, ε) and substitute
into the first equation
Dy hẏ + Dε hε̇ = −h + h2 − y2
or
h0 (y) εy − y3 + xy = −h + h2 − y2 .
Now let
Example 6.2
The Lorenz equation
ẋ = σ(y − x)
ẏ = ρx − y − xz
ż = xy − βz
where σ, ρ and β are positive constants, have a stationary solution at the origin
(x, y, z) = (0, 0, 0). It was shown in Section 2.3 that this stationary solution is
stable if ρ ≤ 0 so we let ρ = 1 + µ, then µ is a bifurcation parameter. We will now
compute the center manifold of the Lorenz equation. First write the system in the
form
ẋ = σ(y − x)
ẏ = (1 + µ)x − y − xz
ż = xy − βz
µ̇ = 0
where we have added µ as a variable. Now the equations can be written in the
form
ż = Az + f (z) (6.1)
where A is the matrix
−σ σ 0 0
1 −1 0 0
A=
0 0 −β 0
0 0 0 0
106 CHAPTER 6. CENTER MANIFOLDS
The eigenvalues of A are 0, −(σ + 1), −β, 0 and the corresponding eigenvectors
are
1 −σ 0 0
1 1 0 0
,
0 0 , 1 , 0 .
0 0 0 1
Now let z = Sy where S is the transformation matrix whose columns are the eigen-
vectors, then
x 1 −σ 0 0 y1 y1 − σy2
y 1 −1 0 0 y2 y1 + y2
=
y3 =
z 0 0 1 0 y3
µ 0 0 0 1 w w
where
1 σ
1+σ 1+σ 0 0
−1 1
0 0
S−1 =
1+σ 1+σ
0 0 1 0
0 0 0 1
and
0 0 0 0
0 −(σ + 1) 0 0
B = S−1 AS =
0 0 −β 0
0 0 0 0
Now σ
y1 − σy2 1+σ (y1 − σy2 )(w − y3 )
1
1+σ (y1 − σy2 )(w − y3 )
y1 + y2
S−1 f
=
y3 (y1 − σy2 )(y1 + y2 )
w 0
107
This equation shows that the Lorenz equations have a two-dimensional center
manifold at the origin and a two dimensional stable manifold. By the Center
Manifold Theorem there exists a hc ∈ Ck such that
hc : Π(U) → E s
in a neighborhood U of the origin, with hc (0, 0) = (0, 0) and Dhc (0, 0) vanishing.
We let
y2 h2 (y1 , w)
= h(y1 , w) =
y3 h3 (y1 , w)
The equation determining the center manifold is
σ
D(y1 ,w) h 1+σ (y1 − σh2 )(w − h3 ) =
−(1 + σ) 0 y2
0 0 −β y3
1
+ 1+σ (y1 − σh2 )(w − h3 )
(y1 − σh2 )(y1 + h3 )
We approximate h by a power series using that the constant and linear terms must
vanish
The above equation for the center manifold can be written as two equations
(6.2)
σ ∂h2 1
(y1 − σh2 )(w − h3 ) = −(1 + σ)h2 + (y1 − σh2 )(w − h3 )
1 + σ ∂y1 1+σ
σ ∂h3
(6.3) (y1 − σh2 )(w − h3 ) = −βh3 + (y1 − σh2 )(y1 + h3 )
1 + σ ∂y1
108 CHAPTER 6. CENTER MANIFOLDS
σ
(2a3 y21 w + b3 y1 w2 ) + O(y1 , w)4
1+σ
(6.5) = −β(a3 y21 + b3 y1 w + c3 w2 + d3 y31 + e3 y21 w + f3 y1 w2 + g3 w3 )
+(y21 w + a3 y31 + b3 y21 w + c3 y1 w2 − σa2 y31 − σb2 y21 w − σc2 y1 w2 )
up to terms of order (y1 , w)4 . Substituting these expressions into the differential
equation for y1 then gives the flow on the center manifold
σ 1 1 σ
ẏ1 = (y1 w − y31 − y21 w − y1 w2 ) + O(y1 , w)4 (6.6)
(1 + σ) β β (1 + σ)2
We will show below that this equation gives a pitchfork bifurcation as µ = w
increases through zero.
Chapter 7
Bifurcation Theory
(7.1) ẋ = Ax + f (x, y)
ẏ = By + g(x, y),
given in Corollary 6.1, and then we can observe what bifurcations take place on
c as the coefficients of the h expansion vary. It turns out that if dim x = 1 or 2
Wloc
one can tell the whole story, these are called the codimension 1 and 2 bifurcations
respectively. The codimension 3, dim x = 3, case is much more complicated and
is still unresolved. For higher dimensional cases, dim x > 3, there is not much
that can be said in general unless symmetries are present so that the bifurcations
take place on lower-dimensional subspaces.
109
110 CHAPTER 7. BIFURCATION THEORY
ẋ = f (x, h(x)).
Namely, since A has only one eigenvalue it must be zero and this says that in the
codimension one case the flow is given by a one-dimensional equation. We are
interested in how the flow changes with parameters in the problem and therefore
consider the one dimensional equation
ẋ = f (x, µ) (7.2)
f (0, 0) = 0,
7.1. CODIMENSION ONE BIFURCATIONS 111
X X
µ µ
This says that locally the bifurcation curve is a parabola centered on the µ-axis.
Moreover, since the stability of the bifurcating solutions (x, µ) are determined by
the linearized equation (7.2),
and
dµ
Dx f (x, µ) = −Dµ f (x, µ)
dx
we get the stability information in Table 7.1.1. Now there are 4 possible bifurca-
tion diagrams which are illustrated on Figures 7.1.1 and 7.1.1.
Table III.1
Dµ f > 0 Dµ f < 0
µ0 >0 Stable Unstable
µ0 <0 Unstable Stable
X X
µ µ
X X
µ µ
x x
µ µ
x x
µ µ
This means that for µ small we can get 3 branches of stationary solutions. Namely,
clear. A differentiation of f (x, µ)/x similar to the one performed in the saddle-
node case above also shows that
dµ d2µ D3x f
= 0 and = − 6= 0,
dx dx2 3Dx Dµ f
for the x± branches. This says that these branches form a parabola along the µ
axis in the µ − x plane, see Figure 7.1.3.
The stability is determined by the linearization of (7.2) about these branches.
For the x = 0 branch we get
Thus the stability of the x = 0 is opposite to that of the x± branches and determined
by the signature of Dxµ f (0, 0), see Figure (7.1.3). The signatures of D3x f and Dxµ f
determine in which µ half-plane we get three branches of stationary solutions,
is called the subcritical case, see Figure 7.1.3. In the former case we get three
branches for µ positive in the latter case we get three branches for µ negative.
Now the canonical example or normal form exhibiting pitchfork bifurcations
is
f (x, µ) = αµx − βx3 , α, β = ±1.
The cases on Figure 7.1.3 correspond to α = 1 = β, α = −1 = β, α = −1, β =
+1 and α = +1, β = −1, respectively.
7.1. CODIMENSION ONE BIFURCATIONS 117
Example 7.1
Recall the equation (6.6) describing the flow on the center manifold of the
Lorenz equations
σ 1 1 σ
ẏ = (yµ − y3 − y2 µ − yµ2 ) + O(y, µ)4 = f (y, µ) (7.14)
(1 + σ) β β (1 + σ)2
We compute the derivatives give us the criteria for a bifurcation and the type of
bifurcation of the stationary solution at the origin (y, µ) = (0, 0).
∂f σ 3 2 σ
= (µ − y2 − yµ − µ2 ) = 0, at (0, 0)
∂y 1 + σ β β (1 + σ)2
∂f σ 1 2σ
= (y − y2 − yµ) = 0, at (0, 0)
∂µ 1 + σ β (1 + σ)2
∂2 f σ 6 2
2
= (− y − µ) = 0, at (0, 0)
∂y 1+σ β β
The first line shows that there is a bifurcation point at the origin. The second line
shows that it is not a saddle-node bifurcation. The third line shows that it is not a
transcritical bifurcation. Next we show that the non-degeneracy conditions for a
pitchfork bifurcation at the origin are satisfied.
∂2 f σ 2 2σ σ
= (1 − y − 2
µ) = 6= 0,
∂y∂µ 1+σ β (1 + σ) 1+σ
∂3 f σ 6
3
= − 6= 0
∂y 1+σ β
Thus the Lorenz equation have a pitchfork bifurcation at the origin and now we
show that it is supercritcal, namely
∂2 f ∂3 f
sign 6= sign 3
∂y∂µ ∂y
More information on the bifurcation theory of ODEs and proofs can be found
in Iooss and Joseph [13] and for PDEs in Chow and Hale [7]. The reduction to
normal forms is performed in Arrowsmith and Place [3].
118 CHAPTER 7. BIFURCATION THEORY
Definition 7.1 Let {{T (t)}, X} be a flow, then we say that a codimension one
hypersurface Γ ⊂ X is a transversal, in a neighborhood U of u ∈ Γ, if every orbit
u(t) = T (t)uo meets Γ again and the vector field u̇, at u, is not tangent to Γ. Γ is
7.3. THE PERIOD DOUBLING BIFURCATION 119
where τ(w) is the first time that the orbit u(t) = T (t)w returns to Γ ∩U, U being
a neighborhood of w in X.
and suppose
f (0, 0) = 0 and Dx f (0, 0) = −1.
This gives rise to a bifurcation which is not possible for one-dimensional flows,
namely a branch of fixed points bifurcates into a periodic orbit. The names comes
from the fact that if the map is the Poincaré map of a flow and the fixed point of the
map corresponds to the periodic orbit of the flow, of period one, then the periodic
orbit of the map corresponds to a periodic orbit of the flow with period two. Now
the implicit function theorem implies that there is a branch of fixed points going
through the origin. We analyze the bifurcation by considering the second iteration
of the map
xm+1 = f 2 (xm , µ)
and consider the flow
ẋ = g(x, µ) = f 2 (x, µ) − x.
The derivatives at the origin are
Dx g = (Dx f )2 (0, 0) − 1 = 0
by chain rule,
Dµ g = Dx f Dµ f (0, 0) = 0,
120 CHAPTER 7. BIFURCATION THEORY
if Dx f = −1, and Dµ f = 0.
In addition
D2x g = D2x f Dx f (Dx f + 1) = 0,
but
D2µx g = 2Dx f Dxµ f (0, 0) + D2x f Dµ f = −2Dxµ f (0, 0) 6= 0,
and
2
D3x g = Dx f D3x f 1 + (Dx f (0, 0))2 + 3 D2x f Dx f (0, 0)
n 2 o
= − 2D3x f (0, 0) + 3 D2x f (0, 0)
6= 0
in general. Thus by the analysis of the pitchfork bifurcation in the previous sec-
tion,
xm+1 = f 2 (xm , µ)
has a pitchfork bifurcation at the origin. This means that if we denote by x+ the
top and x− the bottom pitchfork branch, then
x+ = f (x− , µ)
and
x− = f (x+ , µ).
The reasoning is that the second iterate must be a stable fixed point of f 2 , but there
are only two such fixed points. Moreover,
x− = f (x− , µ)
is impossible because then f would have another branch of fixed points going
through the origin. The following theorem holds,
Dxµ f (0, 0) 6= 0
and 2
2D3x f (0, 0) + 3 D2x f (0, 0) 6= 0,
then the map (7.16) has a period-doubling (also called a flip) bifurcation at the
origin. There exists a stable branch of fixed points for µ < 0 that becomes unstable
for µ > 0, at the origin, and there exists a branch of stable periodic orbits for µ > 0.
7.4. THE HOPF BIFURCATION 121
Using the theory of Poincaré-Birkhoff normal forms, see for example Arrowsmith
and Place [3], the equation (7.17) can be reduced to an equation in polar coordi-
nates,
ṙ = µr + ar3 + O(r5 , µ)
(7.19) θ̇ = ω + br2 + O(r4 , µ).
The bifurcation is controlled by the r equation and it is the pitchfork equation that
we analyzed above. The quantity a is computed from the vector field F = ( f , g),
1
a = ( fxxx + fxyy + gxxy + gyyy )
16
1
(7.20) + ( fxy ( fxx + fyy ) − gxy (gxx + gyy ) − fxx gxx + fyy gyy ) ,
16
see for example Guckenheimer and Holmes [11]. If a < 0 then the bifurcation
is supercritical and we get a branch of stable stationary solution for µ < 0, that
becomes unstable at the origin and throws off a stable periodic orbit. These orbits
form a paraboloid whose center is the branch of unstable stationary solutions for
µ > 0, see Figure 7.4. For µ > 0, the origin in x-space is encircled by a unique
stable periodic orbit whose size and period changes continuously with µ, see [3].
If a > 0 the bifurcation is subcritical and the stability reverses, the unstable
periodic orbit encircles the origin, in x-space, see Figure 7.4, for µ < 0.
122 CHAPTER 7. BIFURCATION THEORY
123
124 CHAPTER 8. THE PERIOD DOUBLING CASCADE
dµ = ρ(x)dx
where S ⊂ [−1, 1] is any subset of the interval. Then we say that the map possesses
an absolutely continuous invariant measures. We shall think about such maps as
being strongly chaotic because they also possess positive Lyapunov exponents and
the associated sensitive dependance on initial conditions.
Now we describe how to construct a map with sensitive dependance on initial
conditions. The method is to let f fall on an unstable fixed point after a few it-
erations of the map. The reason is that f 0 (0) = 0 and this is the greatest stability
one can achieve in any neighborhood. If that neighborhood falls on the neigh-
borhood of an unstable fixed point after a few iterations, then the points in the
8.1. THE QUADRADIC MAP 127
Figure 8.5: The histogram and support of the strange attractor with an absolutely
continuous invariant measure at µ = 2.
130 CHAPTER 8. THE PERIOD DOUBLING CASCADE
f (x) = 1 − 1.544x2 .
The origin is mapped onto the unstable fixed point x in three iterations and nearby
points must disperse. In Figure 8.1 we show the corresponding histogram and the
support of the strange attractor.
Table III.1
µn −µn−1
n µn µn − µn−1 µn+1 −µn
0 .75 .5
1 1.25 .1180989394 4.233738275
2 1.3680989394 .0259472172 4.551506949
3 1.3940461566 .0055850823 4.645807493
4 1.3996312389 .0011975035 4.663938185
5 1.4008287424 .0002565289 4.668103672
6 1.4010852713 .000054943399 4.668966942
7 1.401140214699 .000011767330 4.669147462
8 1.401151982029 .000002520208 4.669190003
9 1.401154502237 .000000539752 4.669196223
20 1.401155041989
The first column shows that the µn values actually converge to a terminal point
µ∞ . The second column shows that the windows of periodicity get smaller and
smaller. The last column shows that the ratio converge to a constant
δ = 4.66920 · · · .
8.2. SCALING BEHAVIOUR 131
Figure 8.7: The histogram and support of the strange attractor with an absolutely
continuous invariant measure, covering part of the interval, at µ = 1.544.
8.2. SCALING BEHAVIOUR 133
Definition 8.1 The periodic points µ0n of period 2n , whose periodic orbits contain
zero, are called superstable.
The name superstable comes from the fact that
n
D2 f (0) = 0,
which gives the strongest linear contraction possible for the period 2n . The super-
stable points satisfy the scaling relationship
(µ00 − µ∞ )
µ0n = + µ∞
δn
We plot in Figure 8.2. the period doubling cascade with − log (µ∞ − µ) on the
vertical axis instead of µ. The result is periodic windows of period 2n . An even
more dramatic result is achieved if we scale the x-axis as well. Figure 8.2. shows
the period doubling sequence with the vertical axis x · |µ∞ − µ|−0.59367 , where the
exponent α = 0.59367 is given by the relationship
λ = δα
between the two universal constants. This gives a periodic bifurcation diagram
with a reflection (about the unstable pitchfork branch) symmetry.
Now we consider the other side of µ∞ . The point marked µµ0 is the point where
f 3 (0) is an unstable fixed point and we get a sequence of values µµn such that at
(µµ0 − µ∞ )
µµn = + µ∞
δn
134 CHAPTER 8. THE PERIOD DOUBLING CASCADE
n
f 3·2 (0) falls on an unstable periodic orbit of period 2n . In the period 3 window f
(3)
will have a periodic orbit of period 3 · 2n−1 at µn , and
(µ3 − µ0∞ )
µ3·2n = n
+ µ0∞
δ
for a new accumulation point µ0∞ . In other words, the scaling holds on both sides
of µ∞ and applies to period doubling of period two on one side and period three
on the other side. Furthermore, the same analysis applies to period tripling,
quadrapling etc. sequences, with new universal constants λ(n) and δ(n), 3, 4, · · · .
Figure 8.2 illustrates the stable and unstable manifold of the renormalization
map in the following theorem.
Theorem 8.1
1
F g(x) = − g ◦ g(−ax),
a
where g(0) = 1 and a is a constant, denotes the period halving map acting on the
space Gε = {g(x) : [−1, 1] → R | g(x) = f (|x|1+ε ), ε ≤ 1}, where f is a bounded
analytic function on [0, 1] satisfying f (0) = 1, ddyf < 0 on [0, 1] and f (1) > −1;
then for ε ≤ 1, F has a unique fixed point gε in Gε , with a negative Schwartzian
derivative. gε is hyperbolic and
Corollary 8.1 For each a ∈ [−1, 1] there exists a unique point ga ∈ W u ⊂ Gε such
that ga (1) = −a. W u intersects the hypersurfaces
Σ1 = {g ∈ Gε | g(1) = 0}
and
e1 = {g ∈ Gε | g3 (1) = −g(1)}
Σ
transversely.
For a proof of the Theorem and the Corollary, see Collet and Eckmann [8].
8.2. SCALING BEHAVIOUR 137
Figure 8.10: The stable and unstable manifolds of the period halving map.
138 CHAPTER 8. THE PERIOD DOUBLING CASCADE
J0 ⊃ J1 ⊃ · · · ,
3. g maps the subintervals of J j onto one another, J j ⊂ g(J j ), such that the
action of g on J j is a cyclic permutation of order 2 j .
(a) lands after finitely many iterates on one of the periodic orbits in 5,
(there are only countably many such orbits), or
(b) converges to J, such that for each j it is eventually contained in J j .
1 n−1
Z
lim ∑ φ(gn (x)) = φ(x)dν,
n→∞ n J
j=0
where x is any point in [−1, 1] whose orbits converge to J and φ is any continuous
function on [−1, 1].
For a proof of the Theorem see Collet and Eckmann [8].
We now define the tools we need for sensitive dependence on initial condi-
tions.
m(I) = m( f −1 (I))
Theorem 8.3 Let c < log 2 be given. Then there exists a set E of positive
Lebesgue measure close to µ = 2, for the map f (x, µ) = 1 − µx2 , such that the
Lyapunov exponent at x = 1 for µ ∈ E is greater than c.
Corollary 8.3 There exists a small positive number α such that
√
| f j (0)| ≥ e−α j
, for all j ≥ 0 (8.2)
The Corollary shows that the contraction rate of the map at the origin is bounded
from below. This is exactly what one needs to prove Jacobson’s Theorem [14]:
Theorem 8.4 If µ ∈ E so that (8.2) holds, then the map x → f (x, µ) has an in-
variant Borel probability measure that is absolutely continuous with respect to
Lebesgue measure.
The period doubling cascade has applications in all branches of science and
engineering. For applications to superconductors see [4], to electrons in quantum
wells driven by lasers see [10] and [1] and to earthquakes see [9].
Example 8.1 The map x → 1 − 2x2 on [−1, 1] is mapped to the map y → 1 − 2|y|
on [−1, 1] by the homeomorphism
r
4 −1 x+1
y = sin ( )−1 (8.3)
π 2
The Lyapunov exponent of the latter map is obviously λ(y) = 2 because Dy f = ±2
depending on whether y is positive or negative. The latter map has the invariant
measure
1 1
dm = dy (8.4)
π (1 − y2 )1/2
This measure is obviously absolutely continuous with respect to Lebesgue mea-
sure, since
1 1
>0
π (1 − y2 )1/2
and m is a probability measure because
Z 1
1 1 1
2 1/2
dy = sin−1 (y)|1−1 = 1
π −1 (1 − y ) π
Exercise 8.1
1. Show that the map (8.3) maps x → 1 − µx2 to y → 1 − 2|y|.
2. Show that the measure (8.4) is invariant under the map y → 1 − 2|y|.
Appendix A
141
142 APPENDIX A. THE HOMOCLINIC ORBITS OF THE PENDULUM
or
ln | sec (z) + tan (z)| = ±(t + t0 )
so exponentiating both sides we get that
This is now an equation that we need to solve for z, to do that we use the trigonom-
etry identity
1 + tan2 (z) = sec2 (z)
or p
sec (z) = 1 + tan2 z.
If we set e±(t+t0 ) = q the equation becomes
q
1 + tan2 (z) + tan (z) = q.
We move tan (z) to the right hand side and square both sides
q
1 + tan2 (z) = q − tan (z)
so
1 + tan2 (z) = q2 − 2q tan (z) + tan2 (z).
Thus
1 = q2 − 2q tan (z)
and solving for tan (z) gives
1
tan (z) = (q − q−1 ).
2
143
However,
1 e±(t+t0 ) − e∓(t+t0 )
(q − q−1 ) = = ± sinh (t + t0 ).
2 2
Thus
tan (z) = ± sinh (t + t0 )
and
x = 2z = ±2 tan−1 (sinh (t + t0 ))
since both sinh and tan are odd functions. Moreover,
cosh (t + t0 )
ẋ = ±2
1 + sinh2 (t + t0 )
1
= ±2 = ±2 sech(t + t0 )
cosh (t + t0 )
1 + sinh2 (t + t0 ) = cosh2 (t + t0 ).
as t −→ ±∞ , we get
(x, y) −→ (±π, 0).
144 APPENDIX A. THE HOMOCLINIC ORBITS OF THE PENDULUM
Bibliography
[4] B.Birnir and R.Grauer. The global attractor of the damped and driven sine-
Gordon equation. Comm. Math. Phys., 162:539–590, 1994.
[6] J. Carr. Applications of Centre Manifold Theory. Springer, New York, 1981.
[7] S.-N. Chow and J. Hale. Methods of Bifurcation Theory. Springer, New
York, 1982.
[8] P. Collet and J.-P. Eckmann. Iterated Maps on the Interval as Dynamical
Systems. Birkhäuser, Boston, 1980. PPhI Progress in Physics.
[10] B. Galdrikian and B.Birnir. Period doubling and strange attractors in quan-
tum wells. Phys. Rev. Lett.., 76(18):3308–11, 1996.
145
146 BIBLIOGRAPHY
[16] J. Moser. Stable and Random Motions in Dynamical Systems. Ann. Math.
Studies 77. Princeton Univ. Press, Princeton NJ, 1973.